You are on page 1of 21

Mechanical Systems and Signal Processing 186 (2023) 109838

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


journal homepage: www.elsevier.com/locate/ymssp

Assessment of various seismic fragility analysis approaches for


structures excited by non-stationary stochastic ground motions
Xu-Yang Cao a , De-Cheng Feng b ,∗, Yue Li c
a College of Civil and Transportation Engineering, Hohai University, Nanjing 210098, China
b Key Laboratory of Concrete and Prestressed Concrete Structures of the Ministry of Education, Southeast University, Nanjing 210096, China
c
Department of Civil and Environmental Engineering, Case Western Reserve University, Cleveland, OH 44106, USA

ARTICLE INFO ABSTRACT

Communicated by M. Beer A probabilistic assessment is performed using different seismic fragility analysis approaches for
structures under non-stationary stochastic ground motions. Four commonly used probabilistic
Keywords:
seismic fragility analysis (PSFA) approaches are adopted, which are least squares regression
Probabilistic performance
Seismic fragility
(LSR), maximum likelihood estimation (MLE), kernel density estimation (KDE) and Monte Carlo
Non-stationary process simulation (MCS). The principles of the four PSFA approaches are first introduced, and the the-
Stochastic response ories of non-stationary stochastic acceleration time series are introduced in light of the spectral
Structural assessment representation of random functions. Then an analytical procedure of different PSFA approaches
is constructed for structures under non-stationary stochastic motions. After that a case study
is carried out to analyze the results of the four PSFA approaches. The demand distributions of
univariate random variable under certain intensity measure levels are discussed, and multiple
fragility curves under different limit states for all the approaches are compared. In general,
the MCS approach is set as the benchmark to verify the effectiveness of other approaches,
but it is computationally consuming. The non-parametric KDE approach is recommended to
estimate the univariate distribution under specified IM levels, because it can reflect the real
distribution characteristics with combined efficiency and accuracy. For the seismic fragility
under different limit states, the LSR approach is suitable for efficient data processing and general
trend assessment, at the expense of accuracy. The MLE approach indicates a better applicability
under a smaller response threshold and slighter damage conditions, while the KDE approach
shows a better applicability under a larger response threshold and severer damage conditions.
The paper compares the accuracy and applicability of various PSFA approaches under multiple
conditions, and provides a basis to link probabilistic hazard and risk analyses for performance
assessment in the future research.

1. Introduction

Earthquake is one of the most destructive natural hazards that causes tremendous social disruption [1,2]. According to the
statistical data, more than ten thousand people suffer from earthquake annually, accompanied with economic losses over billions of
U.S. dollars. Medium or high-intensity magnitude earthquakes in the last half century indicated that the losses are continuing and
expanding with high potential threat [3–5]. Meanwhile, for existing structures, inappropriate design and construction such as weak
stories, short columns, low-quality materials and poor workmanship are the main reasons that lead to severe damage. Moreover, the
aged structures designed by old seismic codes or with non-fortification considerations may not meet the new seismic requirements

∗ Corresponding author.
E-mail addresses: caoxy@hhu.edu.cn (X.-Y. Cao), dcfeng@seu.edu.cn (D.-C. Feng), yue.li10@case.edu (Y. Li).

https://doi.org/10.1016/j.ymssp.2022.109838
Received 5 April 2022; Received in revised form 18 August 2022; Accepted 23 September 2022
Available online 9 October 2022
0888-3270/© 2022 Elsevier Ltd. All rights reserved.
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

after building code improvement. Thus, seismic assessments and performance enhancements of existing structures have always
attracted research interest in structural engineering and attracted researchers all over the world to explore with various analyzing
approaches [6–11].
Seismic analysis approaches can be generally divided into three types: static analysis, response spectrum analysis and dynamic
time history analysis [12,13]. Compared with the static and response spectrum analyses, the time history analysis treats earthquake
ground motions as time series and takes into account the three main elements of ground motions (i.e., amplitude, frequency and
duration) [14]. Worth mentioning is that the occurrence of ground motion has great randomness and complexity, and the structural
responses under different ground motions are quite different. Thus, how to select appropriate records directly affect the result
accuracy in the dynamic time history analysis. Baker [15] proposed conditional mean spectrum approach to select ground motions,
which reflected the mean response spectrum conditioned on the target response at a certain period. The difference with the uniform
hazard spectrum approach was compared, and the practical guidelines in use were summarized correspondingly. Jayaram et al. [16]
adopted an efficient algorithm in computation to filter seismic records, which showed the agreement with the average target response
spectrum. During the period, target distributions were probabilistically generated and an optimization technique was adopted to
improve the accuracy. Bradley [17] proposed a generalized conditional intensity measure for holistic earthquake wave selections,
and the critical steps were concerned with the multivariate distribution construction from the probabilistic seismic hazard analysis.
Katsanos et al. [18] constructed a research review on the alternative earthquake wave selection procedures for structural analysis,
and during the process a series of strategies (e.g., magnitude and distance choice, spectral matching choice, intensity measure choice)
were summarized.
With the development of earthquake theory, researchers find that the three elements (i.e., amplitude, frequency and duration)
to describe the ground motions are not always sufficient to control the engineering structure responses. Some other characteristics
(such as frequency non-stationary characteristics) also have a non-negligible effect on the structural behaviors, especially on the
nonlinear seismic demands under certain circumstances [19–21]. Thus, the non-stationary characteristics of ground motions have
received extensive attention in recent decades. Clough and Penzien [22] assumed the bedrock and site soil as a second-order
linear filter at the same time, and proposed a modified evolution power spectrum model of ground motions considering the low
frequency energy (C-P model). Deodatis [23] improved the C-P model and regraded the site-related parameters as time-varying
parameters, and a power spectrum based algorithm was developed to generate the fully non-stationary stochastic earthquake waves.
Shinozuka and Deodatis [24] reviewed the stochastic ground motion models, and concluded the formulations of white noise, Poisson,
spectral representation and stochastic wave processes. The analytical expressions were given for all models, and the corresponding
superiorities or drawbacks were commented. Conte and Peng [25] proposed a stochastic non-stationary earthquake model to reflect
the intensity and frequency variations of ground motions along with time. The analytical evolutionary power spectral density
was adaptively estimated to fit with the target power spectral density during the process for parameter acquisition. Rezaeian and
Kiureghian [26] developed a white-noise-based non-stationary stochastic function that employed a discretized filtering process, and
the temporal–spectral non-stationary effects of earthquake waves can be separated. The model allowed flexibility in parameter
estimation, and improved the fitness to target spectra for long periods. Cacciola [27] proposed a novel approach to generate
fully non-stationary earthquake waves with spectrum compatibility, and the superposition of two contributions were considered,
which included a fully non-stationary counterpart and a corrective random process. Several examples were given to verify the
model accuracy and efficiency. Sgobba et al. [28] discussed a fully evolutionary approach to define stochastic ground motions by
filtering Gaussian white-noise process, considering the non-stationarity of both motion amplitude and frequency. The validation
study indicated the satisfactory accuracy of the ground motion model in seismic feature representations over a broad response
range.
On the other hand, with the development of the performance-based earthquake engineering (PBEE), the design purpose to
reduce seismic risks and earthquake losses is becoming a research interest in the field of earthquake engineering [29,30]. The
new generation of PBEE decision-making framework has been proposed by the Pacific Earthquake Engineering Research Center
in light of its full probability expressions. The framework takes probabilistic seismic risk analysis as the research objective, takes
probabilistic seismic hazard analysis as the research inducement, and takes probabilistic seismic fragility analysis (PSFA) as the main
research content [31,32]. Multiple mathematical models, analytical assumptions, numerical approaches, and application scenarios
have been adopted for PSFA in recent years [33,34]. Cornell et al. [35] established a formal probabilistic framework for seismic
design and performance assessment of engineering structures, which was on the basis of a probabilistic relationship between
demand and capacity. The lognormal distributions for demand and capacity were assumed and the most direct way to estimate
the fragility parameters was recommended as a least squares regression. Large quantities of dynamic analyses were needed, and the
alternative options such as incremental dynamic analysis [36,37], multiple stripe analysis [38,39] and cloud analysis [40,41] were
developed by researchers. Shinozuka et al. [42] compared the structural fragility between the empirical and analytical results. The
empirical fragility curves were constructed from the collected damage data, while analytical fragility curves were acquired based
on the maximum likelihood method. The confidence intervals were estimated and the median randomness was also interpreted.
Mai et al. [43] adopted the kernel density estimation to establish the fragility curves of a three-storey steel frame. The results
indicated that the approach was not dependent on the intensity measure (IM) of earthquake waves, and the approach alleviated the
need to estimate the prescribed shape parameters of PSFA curves at the same time. Lupoi et al. [44] developed a computationally
efficient method to assess the structural seismic fragility, involving the nonlinear dynamic analyses and probabilistic demand
characterizations, and the fragility curves were also depicted utilizing the Monte Carlo simulation for validation. Two applications
(i.e., steel-concrete box girder and three-dimensional reinforced concrete building) were introduced for further evaluation. Li [45]
proposed the probability density evolution method in view of a physics-based system reliability background. Combined the structural

2
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

failure criteria and physical equations, a general fragility and reliability assessment frame was established. Several applications were
studied and the engineering value was proved. Altieri and Patelli [46] proposed a non-parametric fragility estimation approach,
characterized with the subset classification and failure region mapping. The fragility was generated through coupling the subset
samples to the classification score, and the corresponding case study indicated higher efficiency and accuracy especially for rare
failure domains. In addition, many novel seismic fragility approaches (e.g., multi-linear regression [47–49], spline smoothers [50–
52], neural networks [53–55]) have been extensively explored by researchers and have been well coupled with the PBEE at this
stage.
As mentioned above, the current research mainly focuses on the utilization of statistical approaches into the PSFA field, but the
analysis of their accuracy with each other or applicability under different conditions is quite limited. For example, comparing the
classic parametric PSFA approaches (e.g., power-law regression, maximum likelihood estimation) [56–58] with the non-parametric
PSFA approaches (e.g., kernel density estimation, Monte Carlo simulation, artificial neural networks) [59–61], their deviations with
the accurate fragility results and the preferred strategy under multiple limit states are not clear. Moreover, although there are state-
of-the-art reviews of the seismic fragility methods summarizing the principles and differences [62–65], the present seismic fragility
study mainly focuses on the seismic input via natural selected ground motions or stationary artificial ground motions, while the
considerations about the uncertainty of seismic event and the comparisons under the influence of non-stationary stochastic ground
motions are relatively scarce [66,67].
Thus, the purpose and novelty of this paper is to compare different seismic fragility analysis approaches for structures under non-
stationary stochastic ground motions, and to compare the accuracy and applicability of diverse PSFA approaches under different
conditions. Four popular PSFA approaches are adopted, which are based on least squares regression (LSR), maximum likelihood
estimation (MLE), kernel density estimation (KDE) and Monte Carlo simulation (MCS). The LSR-based and MLE-based approaches are
two most broadly-used classic parametric PSFA approaches, and the KDE-based and MCS-based approaches are two non-parametric
PSFA approaches without the pre-definitions of the fragility forms. The principles of the four PSFA approaches are first introduced,
and then the theories of non-stationary stochastic acceleration time series are introduced in light of the spectral representation of
random functions. After that a case study of reinforced concrete frames is shown to discuss the results of the four PSFA approaches.
The demand distributions of univariate random variable under certain IM levels are discussed, and multiple fragility curves under
different limit states are compared for all the four approaches.

2. Principles of different probabilistic seismic fragility analysis (PSFA) approaches

The principles of four popular PSFA approaches are introduced in this section, and the analyzing procedure of the four PSFA
approaches for structures under non-stationary stochastic ground motions is displayed in Fig. 1. Besides, Fig. 2 presents the brief
schematic views of the four PSFA approaches in this paper. A total of 200 numerical models related to the structural stochastic
parameters and 200 ground motions related to the earthquake stochastic parameters are generated and matched for the first three
approaches (i.e., LSR-based, MLE-based and KDE-based), respectively. Forty intensity levels are considered and 8000 response results
are acquired for analyses. In comparison, for the last approach (i.e., MCS-based), 10 000 numerical models and 10 000 ground
motions are generated and matched. Twenty intensity levels are considered and 200 000 response results are acquired for analyses.
The theory of the non-stationary stochastic acceleration time series and the case study will be introduced in the following sections.
It is worth mentioning that all the results used for PSFA are based on a non-collapsing structure, thus the data over the thresholds
of collapse limit state should be processed [68]. In this paper, the collapse point is adopted as the first point where the secant
slope reaches 10% of the initial slope. This is of great significance, as discussions of whether one method or another is better at
large deformations are quite influenced by this choice [69]. Usually, the appearance of collapse tends to reduce the importance
of fitting non-collapse cases at large drifts, which makes the selection of different fragility approaches at large deformations
rather ambiguous [70,71]. This step is especially important when proposing KDE over MLE as the top contender at higher drifts
(as concluded in Section 5). In addition, in this paper we just give the comparisons and recommendations of different fragility
approaches when the calculation number is sufficient (e.g., 8000 calculations for the first three methods) and the data used for
fitting are the same. We hope to give the accuracy and the applicability of different fragility approaches for various limit states
under this condition. With related to the efficiency-to-accuracy ratio [72], the relevant conclusions are not discussed in this paper
(e.g., when the LSR-based approach is coupled with the cloud analysis, the efficiency of calculation is improved and the number of
calculation can be greatly reduced than the MLE-based, KDE-based and MCS-based approaches).

2.1. Least squares regression (LSR)-based PSFA approach

The LSR-based PSFA approach is the most commonly adopted method in fragility analysis. Through a series of nonlinear dynamic
analyses (e.g., IDA, MSA or cloud analysis), structural engineering demand measures (e.g. maximum inter-stroy drift ratio, maximum
residual inter-stroy drift ratio, peak floor acceleration) can be obtained under different intensity levels. Related to the principles
of this PSFA approach, more references can be found in Kennedy and Ravindra [73], Lallemant et al. [50], and Bakalis and
Vamvatsikos [74]. Based on the lognormal distribution assumptions for both demand and capacity, the seismic fragility can be
expressed in Eq. (1):
[ √ ]
2
𝑃 (𝐷 > 𝐶|𝐼𝑀) = 𝐹 (𝑎, 𝛿𝑐 , 𝜅, 𝜇) = 𝛷 ln(𝑆𝑑|𝐼𝑀 ∕𝛿𝑐 )∕( 𝛽𝑑|𝐼𝑀 + 𝛽𝑐2 ) (1)

3
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 1. The analyzing procedure of the four PSFA approaches for structures under non-stationary stochastic ground motions.

Fig. 2. The brief schematic views of the four PSFA approaches (note: EDP represents engineering demand parameter).

where 𝑃 is the exceeding probability and is the function of intensity measure (𝑎), median structural capacity (𝛿𝑐 ), and two regression
coefficients (𝜅 and 𝜇). 𝑆𝑑|𝐼𝑀 represents the median structural demand, and its relationship with intensity measure commonly
conforms to a power exponential regression, as expressed in Eq. (2). Through converting the coordinates into a logarithmic scale,
the coefficients (𝜅 and 𝜇) can be obtained using the LSR method.

𝑆𝑑|𝐼𝑀 = 𝜅 ⋅ (𝑎)𝜇 , 𝑙𝑛𝑆𝑑|𝐼𝑀 = 𝑙𝑛𝜅 + 𝜇 ⋅ 𝑙𝑛𝑎 (2)


In Eq. (1), 𝛽𝑑|𝐼𝑀 denotes the logarithmic standard deviation of demands, and can be obtained from the LSR method (stable 𝛽𝑑|𝐼𝑀
in Eq. (3)) or from the statistical characteristics for each intensity level (changeable 𝛽𝑑|𝐼𝑀 ). 𝛿𝑐 reflects the randomness of structural
capacity and can be taken as the defined thresholds for different limit states. 𝛽𝑐 denotes the logarithmic standard deviation of

4
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

capacities, and the corresponding value can be adopted as 0.2 for simplification by Jeon et al. [75]. 𝛷[⋅] denotes the standard
normal distribution function.

√𝑁
√∑
𝛽𝑑|𝐼𝑀 = √ (𝑙𝑛𝛿𝑥 − 𝑙𝑛𝑆𝑑|𝐼𝑀 )2 ∕(𝑁 − 2) (3)
𝑥=1

where 𝛿𝑥 represents the individual demand result for each stochastic earthquake accelerogram, and 𝑁 represents the total number
of stochastic earthquake waves.
Worth mentioning herein is that the LSR-based PSFA approach is commonly coupled with the cloud analysis for efficient data
processing. Because the results from the cloud analysis are scattered, the LSR-based approach can be well used to acquire the
parameters for the fragility analysis. On the other hand, the LSR-based PSFA approach is only one of the choices applicable for the
IDA and MSA (not necessary), as the fragility results via the IDA and MSA can also be obtained by other ways. For example, as for
the IDA, the corresponding fragility can also be acquired via transforming into the CDF form, as expressed in Eq. (4).

𝑃 (𝐷 > 𝐶|𝐼𝑀) = 𝑃 (𝐼𝑀𝐶 < 𝐼𝑀|𝐿𝑆) (4)


where 𝐼𝑀𝐶 represents the random variable of intensity for the target limit state 𝐿𝑆. In this study, the LSR-based PSFA approach is
coupled with the IDA for calculation, as illustrated in Section 4.

2.2. Maximum likelihood estimation (MLE)-based PSFA approach

The MLE-based PSFA approach assumes a two-parameter lognormal distribution of fragility curve. Through introducing the MLE
approach, the two parameters of IMs (i.e., median and logarithmic standard deviation) can be obtained. The seismic fragility with
its two parameters can be denoted by Eq. (5), and more reference can be found in Shinozuka et al. [42] and Baker [76]:
[ ]
𝑙𝑛(𝑎) − 𝑙𝑛(𝜆)
𝑃 (𝐷 > 𝐶|𝐼𝑀) = 𝐹 (𝑎, 𝜆, 𝛽) = 𝛷 (5)
𝛽
where 𝑃 is the exceeding probability and is the function of intensity measure (𝑎), median of intensity (𝜆) and logarithmic standard
deviation of intensity (𝛽). 𝛷[⋅] represents the standard normal distribution function. Worth noticing is that the 𝜆 and 𝛽 are only used
to describe the shape characteristics of the fragility curve, but not to represent the distribution characteristics of a specific structural
demand. The likelihood function (𝑄) for the fragility analysis is expressed by Eq. (6):

{ } ∏
𝑁
[ ]𝑚 [ ]1−𝑚𝑖
𝑄(𝜆, 𝛽, 𝑎𝑖 , 𝑖 = 1, 2, … , 𝑁 ) = 𝐹 (𝑎𝑖 , 𝜆, 𝛽) 𝑖 ⋅ 1 − 𝐹 (𝑎𝑖 , 𝜆, 𝛽) (6)
𝑖=1

where 𝑎𝑖 means the 𝑖th intensity value of stochastic motion for calculation, and 𝑁 represents the total number of stochastic ground
motions. 𝑚𝑖 equals 1 or 0 relying on whether or not the structure can sustain the damage state under the intensity 𝑎𝑖 . If the demand
value surpasses the capacity value in the 𝑖th analysis, 𝑚𝑖 = 1, otherwise 𝑚𝑖 = 0. By differentiating Eq. (6), the two parameters 𝜆 and 𝛽
are computed to maximize 𝑄 (Eq. (7)), and the equation can be further transformed into a logarithmic form to simplify calculations as
expressed in Eq. (8). During the analysis, the optimization algorithm can be adopted for Eq. (8) to get the corresponding parameters
(e.g., double-side approximate process within allowable tolerance).
𝜕𝑄(𝜆, 𝛽) 𝜕𝑄(𝜆, 𝛽)
= =0 (7)
𝜕𝜆 𝜕𝛽
𝜕𝑙𝑛𝑄(𝜆, 𝛽) 𝜕𝑙𝑛𝑄(𝜆, 𝛽)
= =0 (8)
𝜕𝜆 𝜕𝛽

2.3. Kernel density estimation (KDE)-based PSFA approach

Through introducing the conditional probability density function 𝑓𝛥 (𝛿|𝐼𝑀 = 𝑎), the KDE-based PSFA approach reformulates the
fragility function as shown in Eq. (9). More references related to the KDE can be found in Mai et al. [43] and Noh et al. [59]:
𝑃 (𝐷 > 𝐶|𝐼𝑀) = 𝐹 (𝑎, 𝛿𝑐 ) = 𝑃 (𝛥 > 𝛿𝑐 |𝐼𝑀 = 𝑎)
+∞ +∞ 𝑓
𝛥,𝐼𝑀 (𝛿, 𝑎)
= 𝑓𝛥 (𝛿|𝐼𝑀 = 𝑎)𝑑𝛿 = 𝑑𝛿
∫ 𝛿𝑐 ∫𝛿𝑐 𝑓𝐼𝑀 (𝑎)
+∞
(9)
𝑓𝛥,𝐼𝑀 (𝛿, 𝑎)𝑑𝛿
∫𝛿𝑐
=
𝑓𝐼𝑀 (𝑎)
where 𝑃 is the exceeding probability. 𝑓𝛥,𝐼𝑀 (𝛿, 𝑎) represents the joint probability distribution of structural demand 𝛥 and intensity
measure 𝐼𝑀. 𝑓𝐼𝑀 (𝑎) represents the marginal probability distribution of intensity measure 𝐼𝑀. If the two PDFs are obtained, the
fragility curve can be calculated by a mere integration. As for the single random variable 𝑋, the KDE of its probability density
function (PDF) is expressed by Eq. (10) [77]:

1 ∑ 𝑥 − 𝑥𝑖
𝑁
𝑓𝑋 (𝑥) = ⋅ 𝐾( ) (10)
𝑁𝑏 𝑖=1 𝑏

5
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

where 𝑥𝑖 denotes each sample of random variable 𝑋, and 𝑁 denotes the number of samples. 𝑏 denotes the bandwidth factor,
and 𝐾(⋅) denotes the kernel function. The classical kernel functions include uniform, triangular and normal functions, and related
researches [43,78] indicate that the type of kernel function shall not affect the estimation quality provided a large number of sample
sets. Thus, a standard normal kernel function is adopted in this paper as presented in Eq. (11).
[ ]
1 ∑
𝑁
1 1 𝑥 − 𝑥𝑖 2
𝑓𝑋 (𝑥) = ⋅ ⋅ 𝑒𝑥𝑝 − ( ) (11)
𝑁𝑏 𝑖=1 (2𝜋)1∕2 2 𝑏

In contrast, the selection of bandwidth 𝑏 is critical because an inappropriate value can result in an undersmoothed or
oversmoothed PDF. The recommended 𝑏 can be chosen as 1.059𝜎 ⋅ 𝑁 −0.2 [77], and 𝜎 denotes the standard deviation of random
variable 𝑋. KDE can be further extended to a multivariate random variable 𝑿 as expressed in Eq. (12):
∑𝑁 ( −1∕2 )
1
𝑓𝑿 (𝒙) = ⋅ 𝐾 𝑩 (𝒙 − 𝒙𝒊 ) (12)
𝑁|𝑩|1∕2
𝑖=1

where 𝑩 represents a definite symmetric bandwidth matrix whose determinant is expressed as |𝑩|. Through applying a multivariate
standard normal kernel, the joint probability distribution is written as Eq. (13):

𝑁 [ ]
1 1 1 −1
𝑓𝑿 (𝒙) = ⋅ ⋅ 𝑒𝑥𝑝 − (𝒙 − 𝒙𝒊 )𝑇 𝑩 (𝒙 − 𝒙𝒊 ) (13)
𝑁|𝑩|1∕2 𝑖=1 (2𝜋)𝑑∕2 2

where 𝒙𝒊 denotes each sample of random variable 𝑿 ∈ 𝑅𝑑 in a 𝑑-dimensional scale. The bandwidth matrix 𝑩 is critical for joint
distribution estimation and can be computed utilizing plug-in or cross-validation estimators, as recommended by Duong [79].
Based on the univariate sample sets of 𝐼𝑀 of all the stochastic ground motions (i.e., 𝐼𝑀𝑖 , 𝑖 = 1, 2, … , 𝑁), the marginal PDF
(𝑓𝐼𝑀 (𝑎)) can be estimated using Eq. (11). Based on the bivariate sample sets of 𝛥 and 𝐼𝑀 of all the stochastic ground motions
(i.e., (𝛥𝑖 , 𝐼𝑀𝑖 ), 𝑖 = 1, 2, … , 𝑁), the joint PDF (𝑓𝛥,𝐼𝑀 (𝛿, 𝑎)) can be estimated using Eq. (13). The final fragility function of Eq. (9) can
be rewritten as Eq. (14):
+∞ ∑𝑁 [ ]
1 𝑎 − 𝐼𝑀𝑖 𝑇 𝑎 − 𝐼𝑀𝑖
𝑒𝑥𝑝 − ( ) ⋅ 𝑩 −1 ⋅ ( ) 𝑑𝛿
𝑏𝐼𝑀 ∫𝛿𝑐 2 𝛿 − 𝛥𝑖 𝛿 − 𝛥𝑖
𝑖=1
𝐹 (𝑎, 𝛿𝑐 ) = ⋅ [ ] (14)
(2𝜋|𝑩|)1∕2 ∑𝑁
1 𝑎 − 𝐼𝑀𝑖 2
𝑒𝑥𝑝 − ( )
𝑖=1
2 𝑏𝐼𝑀

2.4. Monte Carlo simulation (MCS)-based PSFA approach

The MCS-based PSFA approach calculates the unknown characteristics by obtaining mass statistical values according to the
sampling method. The more samples, the closer to the optimal solution. The MCS is commonly set as the benchmark for accuracy
and used to verify the effectiveness of other simulation methods. The MCS-based seismic fragility function can be denoted by Eq. (15):

𝑁𝛿𝑐 (𝑎)
𝑃 (𝐷 > 𝐶|𝐼𝑀) = 𝐹 (𝑎, 𝛿𝑐 ) = (15)
𝑁𝑡𝑜𝑙 (𝑎)
where 𝑃 is the exceeding probability. 𝑁𝑡𝑜𝑙 (𝑎) represents the total samples under the intensity measure 𝑎, and 𝑁𝛿𝑐 (𝑎) represents the
samples that exceed the limit state with the threshold of 𝛿𝑐 under the intensity measure 𝑎.

3. Non-stationary stochastic ground motions

3.1. Difference with stationary stochastic process

With the development of earthquake engineering, researchers realize that ground motions contain strong randomness in space,
time and intensity, and the theory of stochastic processes can be introduced to construct the real earthquake ground motions
for performance assessment [80–82]. Since the 1970s, the spectral representation method to reflect stochastic processes has been
continuously improved, and three types of random models have been proposed for the seismic analysis (i.e., stationary process,
non-stationary process for intensity, and non-stationary process for intensity-frequency). The early stationary process [83] treats the
power spectrum as a constant (e.g., stationary white noise model) and ignores the influence of the site conditions on the acceleration
of ground motions (e.g., site frequency and damping ratio). Although Kanai and Tajimi (K-T model) [84,85], Clough and Penzien (C-P
model) [22] treated the soil layer on bedrock as a filter and improved the power spectrum, the corresponding stochastic processes
are still stationary without the consideration of time variation.
Later researchers found that the earthquake process is from weak stage to strong stage then back to weak stage, and ignoring
the time non-stationarity in intensity or frequency may not reflect the reality of the ground motions. Bolotin [86] multiplied the
stationary process by a deterministic envelope function to obtain the intensity non-stationary ground motion acceleration, which
realized the process of intensity strengthening-weakening along with time. Liu et al. [87] proposed the generalized power spectrum
model and considered the frequency variation with time in the spectrum parameters, which constructed the stochastic process

6
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

combining the intensity-frequency non-stationarity. Thus, in this paper, the non-stationary stochastic acceleration time series are
obtained to reflect the reality of earthquake ground motions and to guarantee the accuracy of calculating results in performance
assessment. Both the intensity non-stationarity and frequency non-stationarity are considered and a modification procedure is
utilized to improve the fitting accuracy with the target acceleration spectrum. The non-stationary stochastic ground motions are
expressed on the basis of the spectral representation of random functions as well as the stochastic process theory [88,89]. The
specific equations are introduced in the following subsection.

3.2. Spectral representation function

Through introducing the Clough–Penzien bilateral evolutionary power spectral density (EPSD) that incorporates the intensity-
frequency non-stationarity, the stochastic seismic motion with zero-mean process and 2𝑁-random variables (𝑋̈ 𝑔 (𝑡)) can be generated
by Eq. (16) [90,91]:
𝑁 √
∑ [ ]
𝑋̈ 𝑔 (𝑡) = 2𝑆𝑋̈ 𝑔 (𝑡, 𝜔𝑘 ) ⋅ 𝛥𝜔 ⋅ cos(𝜔𝑘 𝑡)𝑋𝑘 + sin(𝜔𝑘 𝑡)𝑌𝑘 (16)
𝑘=1

where 𝜔𝑘 = 𝑘𝛥𝜔, and 𝛥𝜔 is interval frequency determined by the truncated frequency (𝜔𝑢 ) and truncated items (𝑁). In this paper,
{ }
𝛥𝜔 is adopted as 0.15 rad/s and 𝑁 is adopted as 1000. 𝑋𝑘 , 𝑌𝑘 (𝑘 = 1, 2, … , 𝑁) represent the standard orthogonal random variables,
which are acquired after a deterministic mapping and rearrangement from the two groups of standard orthogonal random variables
{ } { }
𝑋̄ 𝑛 , 𝑌̄𝑛 (𝑛 = 1, 2, … , 𝑁). This step is critical to avoid sharp amplitude points of and to obtain the ideal acceleration process. 𝑋̄ 𝑛 , 𝑌̄𝑛
can be expressed in the form of two independent random variables 𝛩1 and 𝛩2 , respectively by Eq. (17), thus the dimensions of non-
stationary stochastic earthquake waves can be reduced from 2𝑁 to 2 to greatly improve the efficiency of structural random dynamic
analysis [87].

𝑋̄ 𝑛 = cas(𝑛𝛩1 ), 𝑌̄𝑛 = cas(𝑛𝛩2 ) (17)

where Eq. (17) represents the non-Gaussian orthogonal random form and cas(𝑥) = 𝑠𝑖𝑛(𝑥) + 𝑐𝑜𝑠(𝑥) represents the Hartley orthogonal
function. The fundamental random variables 𝛩1 and 𝛩2 conform to uniform distributions from 0 to 2𝜋. The Clough–Penzien bilateral
EPSD function (𝑆𝑋̈ 𝑔 (𝑡, 𝜔)) is the critical link in the simulation of non-stationary stochastic earthquake waves and can be denoted by
Eq. (18) [23,92]:

2 𝜔4𝑔 (𝑡) + 4𝜉𝑔2 (𝑡)𝜔2𝑔 (𝑡)𝜔2 𝜔4


𝑆𝑋̈ 𝑔 (𝑡, 𝜔) = 𝐹 (𝑡) ⋅ [ ]2 ⋅[ ]2 ⋅ 𝑆0 (𝑡) (18)
𝜔2 − 𝜔2𝑔 (𝑡) + 4𝜉𝑔2 (𝑡)𝜔2𝑔 (𝑡)𝜔2 𝜔2 − 𝜔2𝑓 (𝑡) + 4𝜉𝑓2 (𝑡)𝜔2𝑓 (𝑡)𝜔2

where 𝐹 (𝑡) represents the intensity modulation function and is recommended as the unimodal envelope form proposed by Ou and
Wang (Eq. (19)) [93].
[ ]
𝑡 𝑡 𝑑
𝐹 (𝑡) = ⋅ 𝑒𝑥𝑝(1 − ) (19)
𝑐 𝑐
In Eq. (18), 𝜔𝑔 (𝑡), 𝜉𝑔 (𝑡), 𝜔𝑓 (𝑡) and 𝜉𝑓 (𝑡) reflect the frequency non-stationary characteristics of EPSD, and can be denoted as
Eqs. (20) and (21):
𝑡 𝑡
𝜔𝑔 (𝑡) = 𝜔0 − 𝑎 , 𝜉𝑔 (𝑡) = 𝜉0 + 𝑏 (20)
𝑇 𝑇

𝜔𝑓 (𝑡) = 0.1𝜔𝑔 (𝑡), 𝜉𝑓 (𝑡) = 𝜉𝑔 (𝑡) (21)

In Eq. (18), 𝑆0 (𝑡) represents spectral intensity parameter and can be expressed as Eq. (22) [94]:
𝑎̄2max
𝑆0 (𝑡) = [ ] (22)
𝛾 2 𝜋𝜔𝑔 (𝑡) ⋅ 2𝜉𝑔 (𝑡) + 1∕(2𝜉𝑔 (𝑡))

Among the coefficients from Eqs. (18) to (22), 𝑎 and 𝑏 represent the factors influenced by field classifications and seismic groups,
𝑐 represents the average arrival time of peak ground acceleration (PGA), and 𝑑 represents the shape control index of the intensity
modulation function (𝐹 (𝑡)). 𝜔0 and 𝜉0 represent the primary angular frequency and damping ratio of site soil. 𝑇 represents the
duration of the non-stationary stochastic seismic ground motion. 𝛾 represents the equivalent peak factor and 𝑎̄max represents the
mean value of PGA. The specific values of above-mentioned parameters change with different site classification and design group,
and can be referred from Liu et al. [95]. In this paper, 𝑎 = 5 s−1 , 𝑏 = 0.2, 𝑐 = 6 s, 𝑑 = 2, 𝑇 = 25 s, 𝜔0 = 13.5 s−1 , 𝜉0 = 0.65, and
𝛾 = 2.6 for site classification-III and design group-II. The initial PGA level is 0.1 g, thus the primary 𝑎̄max is adopted as 0.98 m s−2 .
The sample points of 𝛩1 and 𝛩2 for each non-stationary stochastic earthquake wave are obtained according to the LHS technique.
To improve the fitting accuracy between the average stochastic acceleration spectrum and the target acceleration spectrum, a
modification to the EPSD function is performed on the basis of an iterative strategy as expressed in Eq. (23) [96]:

{
𝑆𝑋̈ 𝑔 (𝑡, 𝜔), 0 < 𝜔 ≤ 𝜔𝑐
𝑆𝑋̈ 𝑔 (𝑡, 𝜔)|𝑖+1 = 𝐴𝑅𝑆 𝑇 (𝜔,𝜉)2 (23)
𝑆𝑋̈ 𝑔 (𝑡, 𝜔)|𝑖 ⋅ 𝐴𝑅𝑆 𝑆 (𝜔,𝜉)2 |𝑖
, 𝜔 > 𝜔𝑐

7
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 3. Non-stationary stochastic ground motions.

where 𝑆𝑋̈ 𝑔 (𝑡, 𝜔)|𝑖+1 and 𝑆𝑋̈ 𝑔 (𝑡, 𝜔)|𝑖 represent the (𝑖 + 1)th and 𝑖th iterative times of EPSD function. 𝐴𝑅𝑆 𝑇 (𝜔, 𝜉) represents the
target acceleration spectrum of Chinese seismic design code [97], and 𝐴𝑅𝑆 𝑆 (𝜔, 𝜉)|𝑖 represents the 𝑖th-iteration average acceleration
spectrum of non-stationary stochastic earthquake waves. 𝜉 denotes damping ratio and equals 0.05 herein. 𝜔 = 2𝜋∕𝑇0 , and 𝑇0 denotes
the fundamental period of structures. 𝜔𝑐 denotes the truncated frequency for modification, and the recommended value is 1.57
rad s−1 . The typical non-stationary stochastic acceleration series under PGA = 0.1 g is displayed in Fig. 3(a). The second-order
statistical results between the target and samples (i.e., average acceleration and standard deviation) are displayed in Figs. 3(b) and
3(c). The acceleration response spectrum before and after the 4th modification are displayed in Figs. 3(d) and 3(e), indicating an
ideal matching degree and an allowable discreteness less than 5%.

4. Case study in reinforced concrete frames (RCFs)

To compare the four popular PSFA approaches under non-stationary stochastic earthquake waves, a case study is performed
on two RCFs, which are designed in light of Chinese design code for concrete structures [98]. The design information and initial
dimension of the two RCFs are displayed in Fig. 4. To reflect the uncertainties of structural components [99,100], 36 random
variables related to materials, sizes and loads are selected and the corresponding distribution characteristics are listed in Table 1. To
generate the samples in numerical simulation, the Latin hypercube sampling technique is introduced, which is an efficient stratified
sampling method for multivariate distributions [101,102].
In this study, the OpenSees software is adopted, which is a commonly-used framework developed to simulate the performance of
structural and geotechnical systems subjected to earthquakes [103,104]. Fig. 5 presents the modeling strategy and element selection
for stochastic dynamic analysis. The force-based nonlinear beam–column element is chosen for the frame body, and the Joint2D

8
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 4. Design information and initial dimension of RCF 3 × 6 and RCF 5 × 10.

Fig. 5. Modeling strategy and element selection for RCFs in OpenSees.

element is chosen to reflect the characteristics of the beam–column connections (e.g., concrete shear and steel slip behaviors). The
parameters of the side spring in Joint2D element can be obtained from the unit fiber-section analysis [105], and the parameters
of the middle spring in Joint2D element can be acquired on the basis of modified compression field theory [106]. The detailed
numerical strategy is not elaborated in this paper and can be available from Cao et al. [107,108].
To perform the fragility analysis in the following subsections, four limit states are defined using the damage measure of
transient maximum interstory drift ratio (MIDR). According to FEMA-356 [109], the four commonly-used limit states are adopted as
operational (OP), immediate occupancy (IO), life safety (LS) and collapse prevention (CP). The damage levels increase from slight
yielding under OP to extensive distortion under CP, and the corresponding MIDRs for the moment-resisting RCFs are recommended
as 0.2%, 1%, 2% and 4%, respectively. Worth mentioning is that other damage measures (e.g., peak floor acceleration, residual
interstory drift ratio) can also be adopted to characterize the different limit states of RCFs, and here we only select the MIDR as an
example because it is the mostly-used and broadly-accepted performance index for RCFs in the fragility assessment [110–114].
Using the aforementioned non-stationary stochastic ground motions and deterministic simulation method, the LSR-based PSFA
results for RCF 3 × 6 and RCF 5 × 10 are displayed in Fig. 6. Figs. 6(a) and 6(d) present the nonlinear IDA curves, fractile curves
and lognormal distribution curves. Figs. 6(b) and 6(e) present the LSR results of structural demands and corresponding coefficients
(i.e., 𝜅, 𝜇 and 𝛽𝑑|𝐼𝑀 ). Figs. 6(c) and 6(f) present the LSR-based fragility curves for the adopted four limit states. The median PGAs
with 50% exceeding probability for 3-span-6-story RCF are given as 0.122 g, 0.484 g, 0.882 g and 1.603 g, respectively, and the
median PGAs for 5-span-10-story RCF are given as 0.076 g, 0.365 g, 0.743 g and 1.499 g, respectively.
Using the aforementioned non-stationary stochastic ground motions and deterministic simulation method, the MLE-based PSFA
results for RCF 3 × 6 and RCF 5 × 10 are displayed in Fig. 7. Figs. 7(a) and 7(d) present the variation of 𝜕𝑙𝑛𝑄(𝜆,𝛽)𝜕𝜆
along with the
parameters 𝜆 and 𝛽 for LS limit state. Figs. 7(b) and 7(e) present the variation of 𝜕𝑙𝑛𝑄(𝜆,𝛽)
𝜕𝛽
along with the parameters 𝜆 and 𝛽 for
CP limit state. Figs. 7(c) and 7(f) present the MLE-based fragility curves for the adopted four limit states. The median PGAs with

9
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Table 1
The random variables and distribution parameters.
Random variables Symbol Distribution Mean COV Source
Stochastic motion parameter 𝛩1 Uniform 3.142 (1) 0.577 [115]
Stochastic motion parameter 𝛩2 Uniform 3.142 (1) 0.577 [115]
Concrete bulk density 𝛾 Normal 26.5 (kN/m3 ) 0.0698 [116]
Floor consistent live load 𝐿𝑐1 Gumbel 0.386 (kN/m2 ) 0.464 [117]
Floor temporary live load 𝐿𝑡1 Gumbel 0.356 (kN/m2 ) 0.683 [117]
Roof consistent live load 𝐿𝑐2 Gumbel 0.504 (kN/m2 ) 0.321 [117]
Roof temporary live load 𝐿𝑡2 Gumbel 0.468 (kN/m2 ) 0.538 [117]
Beam span 𝑠𝑏 Normal 6000 (mm) 0.003 [118]
First story height ℎ𝑓 Normal 4500 (mm) 0.003 [118]
Standard story height ℎ𝑎 Normal 3600 (mm) 0.003 [118]
Column height ℎ𝑐 Normal 600 (mm) 0.01 [118]
Beam height ℎ𝑏 Normal 500 (mm) 0.01 [118]
Beam width 𝑤𝑏 Normal 300 (mm) 0.01 [118]
Slab height ℎ𝑠 Normal 120 (mm) 0.01 [119]
Concrete cover height 𝑐ℎ Normal 30 (mm) 0.01 [118]
1-4-floor core concrete compressive strength 𝑓𝑐𝑝,𝑐𝑜𝑟𝑒1 Normal 26.8 (MPa) 0.18 [118]
1-4-floor core concrete tensile strength 𝑓𝑡𝑝,𝑐𝑜𝑟𝑒1 Normal 2.68 (MPa) 0.18 [118]
1-4-floor core concrete peak strain 𝜀𝑐𝑝,𝑐𝑜𝑟𝑒1 Lognormal 0.0025 (1) 0.15 [118]
1-4-floor core concrete ultimate strain 𝜀𝑐𝑢,𝑐𝑜𝑟𝑒1 Lognormal 0.02 (1) 0.52 [119]
1-4-floor cover concrete peak strain 𝜀𝑐𝑝,𝑐𝑜𝑣𝑒𝑟1 Lognormal 0.002 (1) 0.2 [119]
1-4-floor cover concrete ultimate strain 𝜀𝑐𝑢,𝑐𝑜𝑣𝑒𝑟1 Lognormal 0.01 (1) 0.2 [119]
5-10-floor core concrete compressive strength 𝑓𝑐𝑝,𝑐𝑜𝑟𝑒2 Normal 20.0 (MPa) 0.18 [118]
5-10-floor core concrete tensile strength 𝑓𝑡𝑝,𝑐𝑜𝑟𝑒2 Normal 2.0 (MPa) 0.18 [118]
5-10-floor core concrete peak strain 𝜀𝑐𝑝,𝑐𝑜𝑟𝑒2 Lognormal 0.002 (1) 0.15 [118]
5-10-floor core concrete ultimate strain 𝜀𝑐𝑢,𝑐𝑜𝑟𝑒2 Lognormal 0.02 (1) 0.52 [119]
5-10-floor cover concrete peak strain 𝜀𝑐𝑝,𝑐𝑜𝑣𝑒𝑟2 Lognormal 0.002 (1) 0.2 [119]
5-10-floor cover concrete ultimate strain 𝜀𝑐𝑢,𝑐𝑜𝑣𝑒𝑟2 Lognormal 0.01 (1) 0.2 [119]
Rebar diameter in columns 𝑑25 Normal 25 (mm) 0.04 [118]
Rebar diameter in columns 𝑑22 Normal 22 (mm) 0.04 [118]
Rebar diameter in beams 𝑑20 Normal 20 (mm) 0.04 [118]
Rebar diameter in beams 𝑑18 Normal 18 (mm) 0.04 [118]
Rebar diameter in beams 𝑑12 Normal 12 (mm) 0.04 [118]
Rebar placing distance ℎ𝑟 Normal 85 (mm) 0.1 [119]
Rebar yielding strength 𝑓𝑦 Lognormal 378 (MPa) 0.074 [120]
Rebar elastic modulus 𝐸 Lognormal 201 000 (MPa) 0.033 [120]
Rebar hardening ratio 𝑏 Lognormal 0.02 (1) 0.2 [120]
Rebar elastoplastic parameter 𝑅0 Normal 15 (1) 0.1 [104]
Damping ratio 𝜍 Normal 0.05 (1) 0.1 [119]

50% exceeding probability for 3-span-6-story RCF are given as 0.101 g, 0.474 g, 0.837 g and 1.463 g, respectively, and the median
PGAs for 5-span-10-story RCF are given as 0.085 g, 0.383 g, 0.687 g and 1.269 g, respectively.
Using the aforementioned non-stationary stochastic ground motions and deterministic simulation method, the KDE-based
PSFA results are displayed in Fig. 8. The univariate bandwidth 𝑏𝐼𝑀 is calculated as 0.2791 for both 3-span-6-story RCF
and 5-span-10-story RCF, and the bivariate bandwidth matrix 𝑩 is calculated as [0.0200576, 0.0003393; 0.0003393, 0.0000162] and
[0.0209678,0.0004031;0.0004031,0.0000223] for RCF 3 × 6 and RCF 5 × 10, respectively, based on the R package as recommended
by Duong [79]. Figs. 8(a) and 8(d) present the joint PDF along with the PGA and MIDR. Figs. 8(b) and 8(e) present the contour
lines. Figs. 8(c) and 8(f) present the KDE-based fragility curves for the adopted four limit states. Fig. 8(g) presents the marginal PDF
of PGA, Fig. 8(h) presents the integral joint PDF from 𝛿𝑐 to ∞ for RCF 5 × 10, and Fig. 8(i) presents the joint PDF under different
MIDRs for RCF 5 × 10. The median PGAs with 50% exceeding probability for 3-span-6-story RCF are given as 0.042 g, 0.426 g,
0.853 g and 1.531 g, respectively, and the median PGAs for 5-span-10-story RCF are given as 0.033 g, 0.324 g, 0.685 g and 1.278 g,
respectively.
Using the aforementioned non-stationary stochastic ground motions and deterministic simulation method, the MCS-based PSFA
results for RCF 3 × 6 and RCF 5 × 10 are displayed in Fig. 9. Figs. 9(a) and 9(d) present the calculation results (PGA-MIDR pairs) of
200 000 sampling sets. Figs. 9(b) and 9(e) present the fragility surfaces changing with PGA and MIDR. Figs. 9(c) and 9(f) present the
MCS-based fragility curves for the adopted four limit states. The median PGAs with 50% exceeding probability for 3-span-6-story
RCF are given as 0.103 g, 0.468 g, 0.821 g and 1.534 g, respectively, and the median PGAs for 5-span-10-story RCF are given as
0.078 g, 0.367 g, 0.709 g and 1.359 g, respectively.

5. Comparison of PSFA results and discussions

5.1. Demand distribution under certain IM levels

This subsection compares the structural demand (i.e., MIDR) distribution under certain IM levels (i.e., PGA = 0.1 g, 0.2 g and
0.4 g), and both sampling numbers (i.e., N = 200 and 10 000) are analyzed for RCF 3 × 6 and RCF 5 × 10, as displayed in Fig. 10.

10
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 6. LSR-based PSFA approach.

Fig. 7. MLE-based PSFA approach.

The LSR and MLE approaches assume the lognormal distribution, and the results are marked in pink. The KDE and MCS approaches
are marked in red and black, and the histograms are depicted for all the sampling results. Besides, the normal distribution is also

11
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 8. KDE-based PSFA approach.

supplemented in blue for comparison. The solid lines denote the PDF and the dotted lines denote the cumulative distribution function
(CDF).
It can be observed that with PGA increasing from 0.1 g to 0.4 g, the PDFs of all the approaches become flattened with smaller
peak values, and the corresponding CDFs move to right as well indicating larger distribution ranges. Comparing RCF 3 × 6 with RCF
5 × 10, the latter forms a more flexible system, thus the corresponding PDFs and CDFs further show a larger MIDR range and smaller
peak value under the same condition. Provided a large number of sample sets (i.e., N = 10 000), the PDFs and CDFs for KDE and
MCS approaches are almost the same, except for the slight bumps at PDF peak values. MCS approach generally serves as benchmark
reference for various techniques in fragility estimation, thus the non-parametric KDE approach proves the accuracy and authenticity
in a sense. Although the similarities decrease along with the sampling sets reducing (i.e., N = 200), the curve characteristics can
still be reflected by KDE approach and the CDF of KDE approach still shows the best compared to other distributions.
Although the normal and lognormal distributions are broadly adopted in probabilistic assessment to simplify calculations, the
corresponding curves are idealized with predefined shapes. Although the discrepancy to MCS approach reduces as the samples
increase, the corresponding curve shapes still show obvious difference even under a large number of sample sets (i.e., N =
10 000), especially for PDFs. Besides, the fitting degree of RCF 3 × 6 seems better than RCF 5 × 10 under the same condition
because the coupling uncertainties become complicated as the spans/stories increase. As for CDFs, the lognormal distribution covers
larger distribution ranges than other distributions, accompanied with a lower cumulative probability when it approximates to 1.
Meanwhile, it seems that this tendency of lognormal distributions enhances with a larger probability gap from RCF 3 × 6 to RCF
5 × 10 and from PGA = 0.1 g to PGA = 0.4 g. In contrast, the normal distributions illustrate better CDF tendency with lower
discreteness to MCS approach than the lognormal distributions.

12
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 9. MCS-based PSFA approach.

Thus, for the univariate random variable, the real PDF and CDF estimations in probabilistic statistics are recommended to select
the non-parametric KDE approach with combined efficiency and accuracy (e.g., damage measure distribution under a certain limit
state or engineering demand parameter distribution under a certain intensity). Although the non-parametric MCS approach is simple
and accurate in principle, it is computationally consuming with huge amounts of computing time. The normal and lognormal
distributions (e.g., for LSR and MLE approaches) demonstrate discrepancy to MCS approach to a certain degree, but reflect the
similar curve tendency and simplified analysis process as well. Worth noticing is that the PDF and CDF of normal and lognormal
distributions can be expressed in an analytical form, thus can be convolved with probabilistic hazard analysis and risk analysis
to constitute the integrated analytical framework [35], which is significant for analytical assessment in probabilistic theory with
multivariate random variables.

5.2. Seismic fragility with different limit states

Figs. 11 and 12 present the seismic fragility for 3-span-6-story RCF and 5-span-10-story RCF, respectively, in which the four
PSFA approaches and four limit states are all considered. To quantitatively reflect the fitting degree of different PSFA approaches
to MCS approach, a correlation index 𝜂 is introduced as defined in Eq. (24). A smaller 𝜂 indicates a better matching degree of the
approach to MCS.

√ 𝑛
√∑
𝜂 = √ (𝑃𝑓 −𝑡𝑦𝑝1−𝑖 − 𝑃𝑓 −𝑡𝑦𝑝2−𝑖 )2 ∕(𝑛 − 1) (24)
𝑖

where 𝑃𝑓 −𝑡𝑦𝑝1−𝑖 and 𝑃𝑓 −𝑡𝑦𝑝2−𝑖 present the exceeding probability of the 𝑡𝑦𝑝𝑒−1 and 𝑡𝑦𝑝𝑒−2 approaches at the 𝑖th intensity, respectively,
and 𝑛 represents the total number of intensity levels. The specific 𝜂 of all conditions in fragility curves are marked in Figs. 11 and
12 and listed in Table 2. The median PGA with 50% exceeding probability of all the four PSFA approaches and the corresponding
varying percentages are summarized in Table 3.
It can be observed from Figs. 11(a) and 12(a) that the fragility curves present the similar shapes for all the four limit states and
prove the effectiveness of the four PSFA approaches. From RCF 3 × 6 to RCF 5 × 10, the 𝜂 primarily increases under the CP state,
and the interval percentages reach 122%, 38% and 164% for the LSR, MLE and KDE approaches, respectively. This also reflects
the increased uncertainties along with the span or story adding. The specific characteristics of the adopted PSFA approaches are as
follows:

∙ For the LSR approach, the index 𝜂 is between the results of the KDE and MLE approaches for both OP and IO states
(i.e., 𝜂𝐿𝑆𝑅−𝑀𝐶𝑆 = 0.0451/0.0248 for OP/IO states of RCF 3 × 6, and 𝜂𝐿𝑆𝑅−𝑀𝐶𝑆 = 0.0230/0.0215 for OP/IO states of RCF
5 × 10). The index 𝜂 fluctuates up and down for LS and CP states, with the maximum values of 0.0491 (LS) and 0.0649 (CP)

13
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 10. Demand distribution under certain IM levels.

for RCF 3 × 6 and RCF 5 × 10, respectively. However, among all the limit states of both frames, the LSR approach is not an
optimal choice because either KDE or MLE approach presents a smaller deviation and better fitting accuracy. The LSR approach
is suitable for efficient data processing and macro trend assessment, at the expense of result accuracy.
∙ For the MLE approach, the index 𝜂 is smaller than that of the LSR and KDE approaches for all the OP, IO and LS states
of both frames, and the smallest 𝜂 (0.0049) occurs under the OP state of RCF 5 × 10. Take the IO state for instance, the
𝜂𝑀𝐿𝐸−𝑀𝐶𝑆 reaches 0.0082 for RCF 3 × 6 and reaches 0.0186 for RCF 5 × 10. Compared with the LSR and KDE approaches,
the dropping percentages of MLE approach range from 66.9% to 87.0% for RCF 3 × 6, and from 13.5% to 69.9% for RCF
5 × 10, respectively. The index 𝜂 enlarges under the CP state, with the 𝜂𝑀𝐿𝐸−𝑀𝐶𝑆 of 0.0423 and 0.0584 for RCF 3 × 6 and

14
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Table 2
The correlation index (𝜂) of LSR/MLE/KDE to MCS in fragility curves.
Structural type Correlation index OP state IO state LS state CP state
𝜂𝐿𝑆𝑅−𝑀𝐶𝑆 0.0451 0.0248 0.0491 0.0293
RCF 3 × 6 𝜂𝑀𝐿𝐸−𝑀𝐶𝑆 0.0057 0.0082 0.0205 0.0423
𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 0.0522 0.0631 0.0331 0.0121
𝜂𝐿𝑆𝑅−𝑀𝐶𝑆 0.0230 0.0215 0.0399 0.0649
RCF 5 × 10 𝜂𝑀𝐿𝐸−𝑀𝐶𝑆 0.0049 0.0186 0.0194 0.0584
𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 0.0325 0.0619 0.0223 0.0319

RCF 5 × 10, respectively. The results indicate the better applicability of MLE approach under a smaller MIDR threshold and
slighter damage condition.
∙ For the KDE approach, obvious discrepancy is found under the initial limit states (i.e., OP and IO), and the corresponding
𝜂 shows larger deviations than the results of the LSR and MLE approaches (i.e., 𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 = 0.0522/0.0631 for OP/IO states
of RCF 3 × 6, and 𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 = 0.0325/0.0619 for OP/IO states of RCF 5 × 10). The discrepancy drops comparatively under
the LS state, and the 𝜂 is between the results of the LSR and MLE approaches for both frames (i.e., 𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 = 0.0331 and
0.0223, respectively). The 𝜂 further develops under the CP state, with the best correlation for both frames (i.e., 𝜂𝐾𝐷𝐸−𝑀𝐶𝑆 =
0.0121 and 0.0319, respectively). The results demonstrate the better matching degree of KDE approach under a larger MIDR
threshold and severer damage condition.

The same conclusions can be drawn from the median PGA in fragility curves (Table 3). From RCF 3 × 6 to RCF 5 × 10, the
median PGA apparently decreases under all the states and approaches, showing the more structural flexibility of the latter one and
the larger exceeding probability under the same intensity level. The specific characteristics of RCF 3 × 6 and RCF 5 × 10 are as
follows:

∙ For the RCF 3 × 6, the benchmark median PGAs by the MCS approach are obtained as 0.103 g, 0.468 g, 0.821 g and
1.534 g from the OP to CP states, respectively. Under the OP and IO states (Figs. 11(b) and 11(c)), the MLE approach shows
the best accuracy with the varying percentage of 1.49% and 1.28%, and the KDE approach reflects the worst accuracy with
the varying percentage of 59.22% and 8.97%. As for the LS state ((Fig. 11(d)), the MLE approach remains the best with the
gap of 1.95% while the LSR approach shows the largest gap of 7.43%, indicating that the MLE approach is more appropriate
for a slight damage condition. For the CP state ((Fig. 11(e)), the KDE approach presents more ideal effects than the other two
approaches, and the varying percentage is as small as 0.20%. Generally, the varying percentages reduce from the OP to CP
state for the KDE approach, and the results indicate similar tendency as the correlation index 𝜂.
∙ For the RCF 5 × 10, the benchmark median PGAs by MCS approach are acquired as 0.078 g, 0.367 g, 0.709 g and
1.359 g from the OP to CP states, respectively. Under the OP and IO states (Figs. 12(b) and 12(c)), the KDE approach presents
the distinct discrepancy, and the corresponding varying percentages are as high as 57.69% and 11.72%. Under the LS and
CP states (Figs. 12(d) and 12(e)), the discreteness of LSR approach increases and exceeds the other two approaches, with the
varying percentages of 4.80% and 10.30%, respectively. The optimal median PGA (0.687 g) under the LS state is given by MLE
approach with the varying percentage of 3.10%. In contrast, the results of KDE approach demonstrate the minimal deviation
with the median PGA of 1.278 g and varying percentage of 5.96% under the CP state. Both the median PGA and correlation
index 𝜂 illustrate the applicability of KDE approach for a severe damage condition.

Worth mentioning is that the analysis data and comparison results in this paper are based on the non-stationary stochastic
earthquake ground motions as well as the corresponding stochastic process theory [121,122]. On the one hand, compared with the
approach of selecting natural earthquake records from database, the non-stationary stochastic process incorporates more controlling
parameters (e.g., site angular frequency, soil damping ratio, and shape control index) and is more flexible to reflect different/required
characteristics of earthquake ground motions. Besides, a modification procedure can be introduced into the evolutionary power
spectral density function based on an iterative approach, thus the fitting accuracy between the average stochastic acceleration
spectrum and the target acceleration spectrum can be greatly guaranteed with little discrepancy (while the approach of selecting
natural records is limited to the database and only guarantees the primary considered periods) [96,123]. On the other hand,
compared with the early stationary stochastic process, the non-stationary stochastic process treats the evolutionary power spectral
density as a time-varying function (not a constant), and can incorporate the intensity-frequency non-stationarity along with time
variation [88,124]. The non-stationary stochastic process is more complicated and meaningful, thus can better reflect the reality of
the earthquake acceleration series and lead to more accurate results for structural PSFA procedure.

6. Conclusions

Probabilistic assessment of different seismic fragility analysis approaches for structures under non-stationary stochastic earth-
quake waves is performed in this article. The four commonly used PSFA approaches are adopted as LSR-, MLE-, KDE- and MCS-based
techniques, and the case study is performed based on two RCFs. The corresponding accuracy or applicability of diverse PSFA
approaches under different conditions is discussed, from which the following observations may be drawn:

15
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 11. Seismic fragility for RCF 3 × 6 [note: Fig. 11(b)–(e) are zoom-in views from Fig. 11(a)].

Table 3
The median PGA (g) with 50% exceeding probability in fragility curves.
Structural type Fragility approach OP state IO state LS state CP state
LSR 0.122 (18.45%) 0.484 (3.42%) 0.882 (7.43%) 1.603 (4.50%)
RCF 3 × 6 MLE 0.101 (1.49%) 0.474 (1.28%) 0.837 (1.95%) 1.463 (4.63%)
KDE 0.042 (59.22%) 0.426 (8.97%) 0.853 (3.90%) 1.531 (0.20%)
MCS 0.103 0.468 0.821 1.534
LSR 0.076 (2.56%) 0.365 (0.54%) 0.743 (4.80%) 1.499 (10.30%)
RCF 5 × 10 MLE 0.085 (8.97%) 0.383 (4.36%) 0.687 (3.10%) 1.269 (6.62%)
KDE 0.033 (57.69%) 0.324 (11.72%) 0.685 (3.39%) 1.278 (5.96%)
MCS 0.078 0.367 0.709 1.359

Note: Data in parentheses represent the varying percentages compared to MCS.

16
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

Fig. 12. Seismic fragility for RCF 5 × 10 [note: Fig. 12(b)–(e) are zoom-in views from Fig. 12(a)].

(1) The non-stationary stochastic earthquake accelerations are obtained on the basis of the spectral representation of random
functions and stochastic process theory. The dimensions of random variables can be reduced to 2 to greatly improve the
calculation efficiency and a modification to the EPSD function is introduced on the basis of an iterative approach to greatly
improve the fitting accuracy. Compared with the early stationary stochastic process, the non-stationary stochastic process is
more complicated and incorporates the intensity-frequency characteristics along with time variation, thus can better reflect
the reality of the acceleration series of ground motions and lead to more accurate results for structural PSFA procedure.
(2) The LSR and MLE are classic parametric methods for fragility, and the KDE and MCS are non-parametric methods without
predefined function forms. To be specific, the LSR-based PSFA can be expressed as a relation between demand and capacity,
under the lognormal distribution assumption. Linear data regressions are required for demand and IM pairs. The MLE-based
PSFA approach assumes a two-parameter lognormal distribution and the parameters can be solved by differentiating the
likelihood function for a maximum value. The KDE-based PSFA approach is realized by introducing the conditional PDF,
and the joint probability distribution and marginal probability distribution are required for integration. The MCS-based PSFA
approach uses a large sampling, and the approach is commonly used to verify the effectiveness of other simulation methods.

17
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

For the multiple fragility curves under different limit states, the fragility curves of different methods present similarity and
the MCS approach validates the effectiveness of all the other fragility approaches.
(3) For the univariate distribution under certain intensity levels, the real PDF and CDF estimations in probabilistic statistics are
recommended to select the non-parametric KDE approach with combined efficiency and accuracy. Although its similarities
to MCS decrease as number of sampling reduces, the fragility features can still be reflected, with the best CDF compared
to the normal and lognormal distributions. Although the non-parametric MCS approach is simple and accurate in principle,
it is computationally consuming. The normal and lognormal distributions (e.g., for LSR and MLE approaches) demonstrate
discrepancy to the MCS approach to a certain degree, but reflect the similar curve tendency and the simplified analysis process
as well. The lognormal distribution covers larger distribution ranges than the normal, KDE and MCS distributions for CDF,
accompanied with a lower cumulative probability under the same conditions when it approximates to 1, while the normal
distributions illustrate better CDF with smaller discrepancy to the MCS approach.
(4) For the seismic fragility with different limit states, the correlation index 𝜂 of the LSR approach is between the KDE and MLE
approaches (OP, IO), and fluctuates up and down for LS and CP states. However, among all the limit states, the LSR approach
is not an optimal choice because either KDE or MLE approach presents a smaller deviation and better fitting accuracy. The
LSR approach is suitable for efficient data processing and general trend assessment, at the expense of accuracy. For the MLE
approach, the 𝜂 is smaller than that of LSR and KDE approaches (OP, IO and LS). The 𝜂 increases under the CP state, and the
results indicate the better applicability of MLE approach under a smaller MIDR threshold and slighter damage conditions. For
the KDE approach, obvious discrepancy is found under the initial limit states (OP, IO), and the corresponding 𝜂 shows larger
deviations than that of LSR and MLE approaches. The discrepancy decreases comparatively under the LS state, and further
develops under the CP state with the best correlation. The results demonstrate the better matching degree of KDE approach
under a larger MIDR threshold and severer damage conditions.

List of acronyms

Acronym Explanation Acronym Explanation


CDF Cumulative distribution function MCS Monte Carlo simulation
CP Collapse prevention MIDR Maximum interstory drift ratio
EPSD Evolutionary power spectral density MLE Maximum likelihood estimation
IDA Incremental dynamic analysis MSA Multiple stripe analysis
IM Intensity measure OP Operational
IO Immediate occupancy PBEE Performance-based earthquake engineering
KDE Kernel density estimation PDF Probability density function
LHS Latin hypercube sampling PGA Peak ground acceleration
LS Life safety PSFA Probabilistic seismic fragility analysis
LSR Least squares regression RCF Reinforced concrete frame

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared
to influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgments

The first two authors greatly appreciate the National Natural Science Foundation of China (Grant Nos. 52208164 and 52078119),
the Natural Science Foundation of Jiangsu Province (Grant Nos. BK20220984 and BK20211564), China Postdoctoral Science
Foundation (Grant No. 2022M711028), Jiangsu Funding Program for Excellent Postdoctoral Talent (Grant No. 2022ZB187), and
the Zhi-Shan Scholarship from Southeast University.

References

[1] Y. Li, B.R. Ellingwood, Reliability of woodframe residential construction subjected to earthquakes, Struct. Saf. 29 (4) (2007) 294–307.
[2] X.-Y. Cao, G. Wu, D.-C. Feng, Z. Wang, H.-R. Cui, Research on the seismic retrofitting performance of RC frames using SC-PBSPC BRBF substructures,
Earthq. Eng. Struct. Dyn. 49 (8) (2020) 794–816.
[3] R. Han, Y. Li, J. van de Lindt, Seismic risk of base isolated non-ductile reinforced concrete buildings considering uncertainties and mainshock–aftershock
sequences, Struct. Saf. 50 (2014) 39–56.
[4] S. Öncü-Davas, C. Alhan, Reliability of semi-active seismic isolation under near-fault earthquakes, Mech. Syst. Signal Process. 114 (2019) 146–164.

18
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

[5] X.-Y. Cao, D. Shen, D.-C. Feng, C.-L. Wang, Z. Qu, G. Wu, Seismic retrofitting of existing frame buildings through externally attached sub-structures:
State of the art review and future perspectives, J. Build. Eng. 57 (2022) 104904.
[6] C. Alhan, H. Gazi, H. Kurtuluş, Significance of stiffening of high damping rubber bearings on the response of base-isolated buildings under near-fault
earthquakes, Mech. Syst. Signal Process. 79 (2016) 297–313.
[7] F. Kazemi, B. Mohebi, R. Jankowski, Predicting the seismic collapse capacity of adjacent SMRFs retrofitted with fluid viscous dampers in pounding
condition, Mech. Syst. Signal Process. 161 (2021) 107939.
[8] Y. Bai, Y. Ma, Q. Yang, J. Florez-Lopez, X. Li, F. Biondini, Earthquake-induced damage updating for remaining-life assessment of steel frame substructure
systems, Mech. Syst. Signal Process. 159 (2021) 107782.
[9] Q. Zhang, N.-H. Zheng, X.-L. Gu, Z.-Y. Wei, Z. Zhang, Study of the confinement performance and stress-strain response of RC columns with corroded
stirrups, Eng. Struct. 266 (2022) 114476.
[10] J. Shi, X. Wang, Z. Wu, Z. Zhu, Optimization of anchorage and deviator for concrete beams prestressed with external fiber-reinforced polymer tendons,
Compos. Struct. 297 (2022) 115970.
[11] J.-G. Xu, D.-C. Feng, S. Mangalathu, J.-S. Jeon, Data-driven rapid damage evaluation for life-cycle seismic assessment of regional reinforced concrete
bridges, Earthq. Eng. Struct. Dyn. (2022).
[12] S.K. Jang, C.W. Bert, A.G. Striz, Application of differential quadrature to static analysis of structural components, Internat. J. Numer. Methods Engrg. 28
(3) (1989) 561–577.
[13] M. Berrah, E. Kausel, Response spectrum analysis of structures subjected to spatially varying motions, Earthq. Eng. Struct. Dyn. 21 (6) (1992) 461–470.
[14] E. Salajegheh, A. Heidari, Time history dynamic analysis of structures using filter banks and wavelet transforms, Comput. Struct. 83 (1) (2005) 53–68.
[15] J.W. Baker, Conditional mean spectrum: Tool for ground-motion selection, J. Struct. Eng. 137 (3) (2011) 322–331.
[16] N. Jayaram, T. Lin, J.W. Baker, A computationally efficient ground-motion selection algorithm for matching a target response spectrum mean and variance,
Earthq. Spectra 27 (3) (2011) 797–815.
[17] B.A. Bradley, A generalized conditional intensity measure approach and holistic ground-motion selection, Earthq. Eng. Struct. Dyn. 39 (12) (2010)
1321–1342.
[18] E.I. Katsanos, A.G. Sextos, G.D. Manolis, Selection of earthquake ground motion records: A state-of-the-art review from a structural engineering perspective,
Soil Dyn. Earthq. Eng. 30 (4) (2010) 157–169.
[19] J. Chen, F. Kong, Y. Peng, A stochastic harmonic function representation for non-stationary stochastic processes, Mech. Syst. Signal Process. 96 (2017)
31–44.
[20] Z. Zheng, H. Dai, Y. Wang, W. Wang, A sample-based iterative scheme for simulating non-stationary non-Gaussian stochastic processes, Mech. Syst. Signal
Process. 151 (2021) 107420.
[21] F. Kong, R. Han, S. Li, W. He, Non-stationary approximate response of non-linear multi-degree-of-freedom systems subjected to combined periodic and
stochastic excitation, Mech. Syst. Signal Process. 166 (2022) 108420.
[22] R. Clough, J. Penzien, Dynamics of Structures, McGraw-Hill, 1975.
[23] G. Deodatis, Non-stationary stochastic vector processes: seismic ground motion applications, Probab. Eng. Mech. 11 (3) (1996) 149–167.
[24] M. Shinozuka, G. Deodatis, Stochastic process models for earthquake ground motion, Probab. Eng. Mech. 3 (3) (1988) 114–123.
[25] J. Conte, B. Peng, Fully nonstationary analytical earthquake ground-motion model, J. Eng. Mech. 123 (1) (1997) 15–24.
[26] S. Rezaeian, A. Der Kiureghian, A stochastic ground motion model with separable temporal and spectral nonstationarities, Earthq. Eng. Struct. Dyn. 37
(13) (2008) 1565–1584.
[27] P. Cacciola, A stochastic approach for generating spectrum compatible fully nonstationary earthquakes, Comput. Struct. 88 (15–16) (2010) 889–901.
[28] S. Sgobba, P. Stafford, G. Marano, C. Guaragnella, An evolutionary stochastic ground-motion model defined by a seismological scenario and local site
conditions, Soil Dyn. Earthq. Eng. 31 (11) (2011) 1465–1479.
[29] X.-Y. Cao, D.-C. Feng, G. Wu, J.-G. Xu, Probabilistic seismic performance assessment of RC frames retrofitted with external SC-PBSPC BRBF sub-structures,
J. Earthq. Eng. (2021) 1–24.
[30] Q. Zhang, Z.-Y. Wei, X.-L. Gu, Q.-C. Yang, S.-Y. Li, Y.-S. Zhao, Confinement behavior and stress–strain response of square concrete columns strengthened
with carbon textile reinforced concrete (CTRC) composites, Eng. Struct. 266 (2022) 114592.
[31] O. Yazdanpanah, B. Mohebi, F. Kazemi, I. Mansouri, R. Jankowski, Development of fragility curves in adjacent steel moment-resisting frames considering
pounding effects through improved wavelet-based refined damage-sensitive feature, Mech. Syst. Signal Process. 173 (2022) 109038.
[32] F. Mezghani, A.F. del Rincón, P.G. Fernandez, A. de Juan, J. Sanchez-Espiga, F.V. Rueda, Effectiveness study of wire mesh vibration damper for sensitive
equipment protection from seismic events, Mech. Syst. Signal Process. 164 (2022) 108160.
[33] Y. Li, B.R. Ellingwood, Framework for multihazard risk assessment and mitigation for wood-frame residential construction, J. Struct. Eng. 135 (2) (2009)
159–168.
[34] S. Bjarnadottir, Y. Li, M.G. Stewart, A probabilistic-based framework for impact and adaptation assessment of climate change on hurricane damage risks
and costs, Struct. Saf. 33 (3) (2011) 173–185.
[35] C.A. Cornell, F. Jalayer, R.O. Hamburger, D.A. Foutch, Probabilistic basis for 2000 SAC federal emergency management agency steel moment frame
guidelines, J. Struct. Eng. 128 (4) (2002) 526–533.
[36] D. Vamvatsikos, C.A. Cornell, Incremental dynamic analysis, Earthq. Eng. Struct. Dyn. 31 (3) (2002) 491–514.
[37] D. Vamvatsikos, C.A. Cornell, Applied incremental dynamic analysis, Earthq. Spectra 20 (2) (2004) 523–553.
[38] A. Singhal, A.S. Kiremidjian, Method for probabilistic evaluation of seismic structural damage, J. Struct. Eng. 122 (12) (1996) 1459–1467.
[39] F. Jalayer, C. Cornell, Alternative non-linear demand estimation methods for probability-based seismic assessments, Earthq. Eng. Struct. Dyn. 38 (8)
(2009) 951–972.
[40] F. Jalayer, R. De Risi, G. Manfredi, Bayesian Cloud Analysis: efficient structural fragility assessment using linear regression, Bull. Earthq. Eng. 13 (4)
(2015) 1183–1203.
[41] A. Miano, F. Jalayer, H. Ebrahimian, A. Prota, Cloud to IDA: Efficient fragility assessment with limited scaling, Earthq. Eng. Struct. Dyn. 47 (5) (2018)
1124–1147.
[42] M. Shinozuka, M.Q. Feng, J. Lee, T. Naganuma, Statistical analysis of fragility curves, J. Eng. Mech. 126 (12) (2000) 1224–1231.
[43] C. Mai, K. Konakli, B. Sudret, Seismic fragility curves for structures using non-parametric representations, Front. Struct. Civ. Eng. 11 (2) (2017) 169–186.
[44] G. Lupoi, P. Franchin, A. Lupoi, P.E. Pinto, Seismic fragility analysis of structural systems, J. Eng. Mech. 132 (4) (2006) 385–395.
[45] J. Li, A PDEM-based perspective to engineering reliability: From structures to lifeline networks, Front. Struct. Civ. Eng. (2020) 1–10.
[46] D. Altieri, E. Patelli, An efficient approach for computing analytical non-parametric fragility curves, Struct. Saf. 85 (2020) 101956.
[47] I. Iervolino, G. Baltzopoulos, D. Vamvatsikos, R. Baraschino, SPO2frag v1. 0: software for PUSHOVER-BASED derivation of seismic fragility curves, in:
Proceedings of the VII European Congress on Computational Methods in Applied Sciences and Engineering, ECCOMAS, Crete Island, Greece, 2016, pp.
5–10.
[48] A. Marotta, D. Liberatore, L. Sorrentino, Vulnerability assessment of Italian unreinforced masonry churches using multi-linear regression models, in: 12th
International Conference on Structural Analysis of Historical Constructions (SAHC), 2021.
[49] R. Gentile, C. Galasso, Simplicity versus accuracy trade-off in estimating seismic fragility of existing reinforced concrete buildings, Soil Dyn. Earthq. Eng.
144 (2021) 106678.

19
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

[50] D. Lallemant, A. Kiremidjian, H. Burton, Statistical procedures for developing earthquake damage fragility curves, Earthq. Eng. Struct. Dyn. 44 (9) (2015)
1373–1389.
[51] G. Baltzopoulos, R. Baraschino, I. Iervolino, D. Vamvatsikos, Dynamic analysis of single-degree-of-freedom systems (DYANAS): A graphical user interface
for OpenSees, Eng. Struct. 177 (2018) 395–408.
[52] C.G. Lachanas, D. Vamvatsikos, Rocking incremental dynamic analysis, Earthq. Eng. Struct. Dyn. 51 (3) (2022) 688–703.
[53] N.D. Lagaros, M. Fragiadakis, Fragility assessment of steel frames using neural networks, Earthq. Spectra 23 (4) (2007) 735–752.
[54] S. Mangalathu, G. Heo, J.-S. Jeon, Artificial neural network based multi-dimensional fragility development of skewed concrete bridge classes, Eng. Struct.
162 (2018) 166–176.
[55] Z. Wang, N. Pedroni, I. Zentner, E. Zio, Seismic fragility analysis with artificial neural networks: Application to nuclear power plant equipment, Eng.
Struct. 162 (2018) 213–225.
[56] S.L. Dimova, K. Hirata, Simplified seismic fragility analysis of structures with two types of friction devices, Earthq. Eng. Struct. Dyn. 29 (8) (2000)
1153–1175.
[57] M.B.D. Hueste, J.-W. Bai, Seismic retrofit of a reinforced concrete flat-slab structure: Part II—Seismic fragility analysis, Eng. Struct. 29 (6) (2007)
1178–1188.
[58] R. Gentile, C. Galasso, Gaussian process regression for seismic fragility assessment of building portfolios, Struct. Saf. 87 (2020) 101980.
[59] H.Y. Noh, D. Lallemant, A.S. Kiremidjian, Development of empirical and analytical fragility functions using kernel smoothing methods, Earthq. Eng.
Struct. Dyn. 44 (8) (2015) 1163–1180.
[60] S. Mangalathu, J.-S. Jeon, J.E. Padgett, R. DesRoches, ANCOVA-based grouping of bridge classes for seismic fragility assessment, Eng. Struct. 123 (2016)
379–394.
[61] J. Kiani, C. Camp, S. Pezeshk, On the application of machine learning techniques to derive seismic fragility curves, Comput. Struct. 218 (2019) 108–122.
[62] A. Muntasir Billah, M. Shahria Alam, Seismic fragility assessment of highway bridges: a state-of-the-art review, Struct. Infrastruct. Eng. 11 (6) (2015)
804–832.
[63] M.A. Hariri-Ardebili, V.E. Saouma, Seismic fragility analysis of concrete dams: A state-of-the-art review, Eng. Struct. 128 (2016) 374–399.
[64] I. Zentner, M. Gündel, N. Bonfils, Fragility analysis methods: Review of existing approaches and application, Nucl. Eng. Des. 323 (2017) 245–258.
[65] S. Misra, J.E. Padgett, Seismic fragility of railway bridge classes: methods, models, and comparison with the state of the art, J. Bridge Eng. 24 (12)
(2019) 04019116.
[66] I. Iervolino, Assessing uncertainty in estimation of seismic response for PBEE, Earthq. Eng. Struct. Dyn. 46 (10) (2017) 1711–1723.
[67] I. Iervolino, Estimation uncertainty for some common seismic fragility curve fitting methods, Soil Dyn. Earthq. Eng. 152 (2022) 107068.
[68] S. Mangalathu, J.-S. Jeon, R. DesRoches, Critical uncertainty parameters influencing seismic performance of bridges using lasso regression, Earthq. Eng.
Struct. Dyn. 47 (3) (2018) 784–801.
[69] E. Brunesi, R. Nascimbene, F. Parisi, N. Augenti, Progressive collapse fragility of reinforced concrete framed structures through incremental dynamic
analysis, Eng. Struct. 104 (2015) 65–79.
[70] A. Babič, M. Dolšek, Seismic fragility functions of industrial precast building classes, Eng. Struct. 118 (2016) 357–370.
[71] W.-C. Chuang, S.M. Spence, A performance-based design framework for the integrated collapse and non-collapse assessment of wind excited buildings,
Eng. Struct. 150 (2017) 746–758.
[72] J. Park, P. Towashiraporn, J.I. Craig, B.J. Goodno, Seismic fragility analysis of low-rise unreinforced masonry structures, Eng. Struct. 31 (1) (2009)
125–137.
[73] R. Kennedy, M. Ravindra, Seismic fragilities for nuclear power plant risk studies, Nucl. Eng. Des. 79 (1) (1984) 47–68.
[74] K. Bakalis, D. Vamvatsikos, Seismic fragility functions via nonlinear response history analysis, J. Struct. Eng. 144 (10) (2018) 04018181.
[75] J. Jeon, R. DesRoches, I. Brilakis, L. Lowes, Modeling and fragility analysis of non-ductile reinforced concrete buildings in low-to-moderate seismic zones,
in: Structures Congress 2012, 2012, pp. 2199–2210.
[76] J.W. Baker, Efficient analytical fragility function fitting using dynamic structural analysis, Earthq. Spectra 31 (1) (2015) 579–599.
[77] M.P. Wand, M.C. Jones, Kernel Smoothing, CRC Press, 1994.
[78] C.V. Mai, B. Sudret, K.R. Mackie, B. Stojadinovic, K. Konakli, Non-parametric fragility curves for bridges using recorded ground motions, in: Proceedings
of the 9th European Conference on Structural Dynamic (EURODYN 2014), Faculty of Engineering of University of Porto, 2014, pp. 2831–2838.
[79] T. Duong, et al., Ks: Kernel density estimation and kernel discriminant analysis for multivariate data in r, J. Stat. Softw. 21 (7) (2007) 1–16.
[80] M. Ping, X. Han, C. Jiang, X. Xiao, A time-variant uncertainty propagation analysis method based on a new technique for simulating non-Gaussian
stochastic processes, Mech. Syst. Signal Process. 150 (2021) 107299.
[81] C. Dang, P. Wei, M. Beer, An approach to evaluation of EVD and small failure probabilities of uncertain nonlinear structures under stochastic seismic
excitations, Mech. Syst. Signal Process. 152 (2021) 107468.
[82] D. Jerez, H. Jensen, M. Beer, Reliability-based design optimization of structural systems under stochastic excitation: An overview, Mech. Syst. Signal
Process. 166 (2022) 108397.
[83] G.W. Housner, Characteristics of strong-motion earthquakes, Bull. Seismol. Soc. Am. 37 (1) (1947) 19–31.
[84] K. Kanai, Semi-empirical formula for the seismic characteristics of the ground, Bull. Earthq. Res. Inst. Univ. Tokyo 35 (2) (1957) 309–325.
[85] H. Tajimi, A statistical method of determing the maximum response of a building structure during an earthquake, in: Proc. 2nd World Conf. Earthq.
Eng., 1960, pp. 781–797.
[86] V. Bolotin, Statistical Methods in the Non-Linear Theory of Elastic Shells, Tech. Rep., GENERAL DYNAMICS/ASTRONAUTICS SAN DIEGO CALIF, 1960.
[87] Z. Liu, W. Liu, Y. Peng, Random function based spectral representation of stationary and non-stationary stochastic processes, Probab. Eng. Mech. 45
(2016) 115–126.
[88] Z. Liu, Z. Liu, Random function representation of stationary stochastic vector processes for probability density evolution analysis of wind-induced structures,
Mech. Syst. Signal Process. 106 (2018) 511–525.
[89] F. Kong, Y. Zhang, Y. Zhang, Non-stationary response power spectrum determination of linear/non-linear systems endowed with fractional derivative
elements via harmonic wavelet, Mech. Syst. Signal Process. 162 (2022) 108024.
[90] M. Shinozuka, G. Deodatis, Simulation of stochastic processes by spectral representation, 1991.
[91] M. Shinozuka, G. Deodatis, Simulation of multi-dimensional Gaussian stochastic fields by spectral representation, 1996.
[92] P. Cacciola, G. Deodatis, A method for generating fully non-stationary and spectrum-compatible ground motion vector processes, Soil Dyn. Earthq. Eng.
31 (3) (2011) 351–360.
[93] J. Ou, G. Wang, Random Vibration of Structures (in Chinese), Higher Education Press, 1998.
[94] H. Seya, M.E. Talbott, H.H. Hwang, Probabilistic seismic analysis of a steel frame structure, Probab. Eng. Mech. 8 (2) (1993) 127–136.
[95] Z. Liu, B. Zeng, L. Wu, Spectral representation of non-stationary ground motion process simulation: Random function method (in Chinese), J. Vib. Eng.
28 (3) (2015) 411–417.
[96] R. Pang, B. Xu, D. Zou, X. Kong, Stochastic seismic performance assessment of high CFRDs based on generalized probability density evolution method,
Comput. Geotech. 97 (2018) 233–245.
[97] MHURD-PRC (Ministry of Housing and Urban-Rural Development of the People’s Republic of China), Code for Seismic Design of Buildings GB50011,
Beijing, 2010, (in Chinese).

20
X.-Y. Cao et al. Mechanical Systems and Signal Processing 186 (2023) 109838

[98] MHURD-PRC (Ministry of Housing and Urban-Rural Development of the People’s Republic of China), Code for Design of Concrete Structures GB50010,
Beijing, 2010, (in Chinese).
[99] F. Lorenzoni, F. Casarin, M. Caldon, K. Islami, C. Modena, Uncertainty quantification in structural health monitoring: Applications on cultural heritage
buildings, Mech. Syst. Signal Process. 66 (2016) 268–281.
[100] M. Guedri, S. Cogan, N. Bouhaddi, Robustness of structural reliability analyses to epistemic uncertainties, Mech. Syst. Signal Process. 28 (2012) 458–469.
[101] J.C. Helton, F.J. Davis, Latin hypercube sampling and the propagation of uncertainty in analyses of complex systems, Reliab. Eng. Syst. Saf. 81 (1) (2003)
23–69.
[102] X. Zhai, C.-W. Fei, Y.-S. Choy, J.-J. Wang, A stochastic model updating strategy-based improved response surface model and advanced Monte Carlo
simulation, Mech. Syst. Signal Process. 82 (2017) 323–338.
[103] F. McKenna, G.L. Fenves, M.H. Scott, et al., Open System for Earthquake Engineering Simulation, University of California, Berkeley, CA, 2000.
[104] S. Mazzoni, F. McKenna, M.H. Scott, G.L. Fenves, et al., OpenSees Command Language Manual, Pacific Earthquake Engineering Research (PEER) Center,
Berkeley, CA, 2006.
[105] X.-Y. Cao, D.-C. Feng, G. Wu, Seismic performance upgrade of RC frame buildings using precast bolt-connected steel-plate reinforced concrete frame-braces,
Eng. Struct. 195 (2019) 382–399.
[106] F.J. Vecchio, M.P. Collins, The modified compression-field theory for reinforced concrete elements subjected to shear, ACI J. 83 (2) (1986) 219–231.
[107] X.-Y. Cao, G. Wu, J.-W.W. Ju, Seismic performance improvement of existing RCFs using external PT-PBSPC frame sub-structures: Experimental verification
and numerical investigation, J. Build. Eng. 46 (2022) 103649.
[108] X.-Y. Cao, C.-Z. Xiong, D.-C. Feng, G. Wu, Dynamic and probabilistic seismic performance assessment of precast prestressed reinforced concrete frames
incorporating slab influence through three-dimensional spatial model, Bull. Earthq. Eng. (2022) 1–35.
[109] FEMA (Federal Emergency Management Agency), Commentary for the Seismic Rehabilitation of Buildings (FEMA-356), Washington, DC, 2000.
[110] X.-Y. Cao, D.-C. Feng, Z. Wang, G. Wu, Parametric investigation of the assembled bolt-connected buckling-restrained brace and performance evaluation
of its application into structural retrofit, J. Build. Eng. (2022) 103988.
[111] X.-Y. Cao, D.-C. Feng, G. Wu, Z. Wang, Experimental and theoretical investigations of the existing reinforced concrete frames retrofitted with the novel
external SC-PBSPC BRBF sub-structures, Eng. Struct. 256 (2022) 113982.
[112] D.C. Feng, S.C. Xie, Y. Li, et al., Time-dependent reliability-based redundancy assessment of deteriorated rc structures against progressive collapse
considering corrosion effect, Struct. Saf. 89 (2021) 102061.
[113] D. Feng, X. Ren, J. Li, Stochastic damage hysteretic model for concrete based on micromechanical approach, Int. J. Non-Linear Mech. 83 (2016) 15–25.
[114] Z.P. Chen, S. Zhu, H. Yu, et al., Development of novel SMA-based D-type self-centering eccentrically braced frames, Eng. Struct. 260 (2022) 114228.
[115] J. Xu, D.-C. Feng, Seismic response analysis of nonlinear structures with uncertain parameters under stochastic ground motions, Soil Dyn. Earthq. Eng.
111 (2018) 149–159.
[116] X.-Y. Cao, D.-C. Feng, G. Wu, Pushover-based probabilistic seismic capacity assessment of RCFs retrofitted with PBSPC BRBF sub-structures, Eng. Struct.
234 (2021) 111919.
[117] CABR (China Academy of Building Research), Unified Standards for Building Structure Design (GBJ68-84), Beijing, 1984, (in Chinese).
[118] D.-C. Feng, S.-C. Xie, J. Xu, K. Qian, Robustness quantification of reinforced concrete structures subjected to progressive collapse via the probability
density evolution method, Eng. Struct. 202 (2020) 109877.
[119] X. Yu, Probabilistic Seismic Fragility and Risk Analysis of Reinforced Concrete Frame Structures (Ph.D Thesis), of Harbin Institute of Technology, China.
[120] M. Barbato, Q. Gu, J. Conte, Probabilistic push-over analysis of structural and soil-structure systems, J. Struct. Eng. 136 (11) (2010) 1330–1341.
[121] Z. Huang, Y.-L. Xu, T. Tao, Multi-taper S-transform method for evolutionary spectrum estimation, Mech. Syst. Signal Process. 168 (2022) 108667.
[122] J. Xian, C. Su, Stochastic optimization of uncertain viscous dampers for energy-dissipation structures under random seismic excitations, Mech. Syst. Signal
Process. 164 (2022) 108208.
[123] R. Pang, B. Xu, X. Kong, D. Zou, Y. Zhou, Seismic reliability assessment of earth-rockfill dam slopes considering strain-softening of rockfill based on
generalized probability density evolution method, Soil Dyn. Earthq. Eng. 107 (2018) 96–107.
[124] D. De Domenico, G. Quaranta, G. Ricciardi, W. Lacarbonara, Optimum design of tuned mass damper with pinched hysteresis under nonstationary stochastic
seismic ground motion, Mech. Syst. Signal Process. 170 (2022) 108745.

21

You might also like